ArticlePDF Available

Structural comparison of differently glycosylated forms of acid--glucosidase, the defective enzyme in Gaucher disease

Authors:

Abstract and Figures

Gaucher disease is caused by mutations in the gene encoding acid-beta-glucosidase. A recombinant form of this enzyme, Cerezyme, is used to treat Gaucher disease patients by ;enzyme-replacement therapy'. Crystals of Cerezyme after its partial deglycosylation were obtained earlier and the structure was solved to 2.0 A resolution [Dvir et al. (2003), EMBO Rep. 4, 704-709]. The crystal structure of unmodified Cerezyme is now reported, in which a substantial number of sugar residues bound to three asparagines via N-glycosylation could be visualized. The structure of intact fully glycosylated Cerezyme is virtually identical to that of the partially deglycosylated enzyme. However, the three loops at the entrance to the active site, which were previously observed in alternative conformations, display additional variability in their structures. Comparison of the structure of acid-beta-glucosidase with that of xylanase, a bacterial enzyme from a closely related protein family, demonstrates a close correspondence between the active-site residues of the two enzymes.
Content may be subject to copyright.
electronic reprint
Acta Crystallographica Section D
Biological
Crystallography
ISSN 0907-4449
Editors: E. N. Baker and Z. Dauter
Structural comparison of differently glycosylated forms of
acid-
¬
-glucosidase, the defective enzyme in Gaucher disease
Boris Brumshtein, Mark R. Wormald, Israel Silman, Anthony H. Futerman and Joel
L. Sussman
Copyright © International Union of Crystallography
Author(s) of this paper may load this reprint on their own web site provided that this cover page is retained. Republication of this article or its
storage in electronic databases or the like is not permitted without prior permission in writing from the IUCr.
Acta Cryst.
(2006). D62, 1458–1465 Brumshtein
et al.
¯
Acid-
¬
-glucosidase
research papers
1458 doi:10.1107/S0907444906038303 Acta Cryst. (2006). D62, 1458–1465
Acta Crystallographica Section D
Biological
Crystallography
ISSN 0907-4449
Structural comparison of differently glycosylated
forms of acid-b-glucosidase, the defective enzyme
in Gaucher disease
Boris Brumshtein,
a
Mark R.
Wormald,
b
Israel Silman,
c
Anthony H. Futerman
d
and
Joel L. Sussman
a
*
a
Department of Structural Biology, Weizmann
Institute of Science, Israel,
b
Oxford
Glycobiology Institute, Department of
Biochemistry, University of Oxford,
Oxford OX1 3QU, England,
c
Department of
Neurobiology, Weizmann Institute of Science,
Israel, and
d
Department of Biological Chemistry,
Weizmann Institute of Science, Israel
Correspondence e-mail:
joel.sussman@weizmann.ac.il
#2006 International Union of Crystallography
Printed in Denmark all rights reserved
Gaucher disease is caused by mutations in the gene encoding
acid--glucosidase. A recombinant form of this enzyme,
Cerezyme
1
, is used to treat Gaucher disease patients by
‘enzyme-replacement therapy’. Crystals of Cerezyme
1
after
its partial deglycosylation were obtained earlier and the
structure was solved to 2.0 A
˚resolution [Dvir et al. (2003),
EMBO Rep. 4, 704–709]. The crystal structure of unmodified
Cerezyme
1
is now reported, in which a substantial number of
sugar residues bound to three asparagines via N-glycosylation
could be visualized. The structure of intact fully glycosylated
Cerezyme
1
is virtually identical to that of the partially
deglycosylated enzyme. However, the three loops at the
entrance to the active site, which were previously observed in
alternative conformations, display additional variability in
their structures. Comparison of the structure of acid--gluco-
sidase with that of xylanase, a bacterial enzyme from a closely
related protein family, demonstrates a close correspondence
between the active-site residues of the two enzymes.
Received 4 July 2006
Accepted 19 September 2006
PDB Reference: acid-
-glucosidase, 2j25, r2j25sf.
1. Introduction
Gaucher disease, the most common lysosomal storage
disorder (Futerman & van Meer, 2004), is caused by mutations
in the gene encoding acid--glucosidase (glucocerebrosidase,
GlcCerase; EC 3.2.1.45; Beutler & Grabowski, 2001), resulting
in intracellular accumulation of glucosylceramide (GlcCer).
Fig. 1 shows the reaction catalyzed by GlcCerase. GlcCerase is
a 497 amino-acid residue enzyme with a molecular weight of
62 kDa (Horowitz et al., 1989; Grabowski et al., 1990).
Mutations in GlcCerase diminish activity either by reducing
catalytic activity or by reducing its lysosomal concentration. In
the former case the mutations affect turnover number,
substrate affinity and/or activator binding and in the latter
they compromise folding in the endoplasmic reticulum,
resulting in proteasomal degradation of the protein (Sawkar et
al., 2005). GlcCerase is activated in lysosomes by saposin C
(SapC; Bruhn, 2005; Vaccaro et al., 1997), although neither the
mechanism of activation nor the precise role of SapC are well
understood (Bruhn, 2005).
We previously determined the three-dimensional structure
of Cerezyme
1
(Premkumar et al., 2005; Dvir et al., 2003), a
recombinant form of GlcCerase that is used in enzyme-
replacement therapy (ERT) for Gaucher disease patients
(Jmoudiak & Futerman, 2005). The protein consists of three
non-contiguous domains, with the catalytic site located in
domain III (residues 76–381 and 416–430), a (/)
8
(TIM)
barrel. Although the function of the two non-catalytic
domains is unknown, mutations that cause Gaucher disease
are found in all three domains.
electronic reprint
There are five putative glycosylation sites in GlcCerase, four
of which are believed to be occupied (Asn19, Asn59, Asn146
and Asn270; Grace et al., 1994). In order to target GlcCerase
to macrophages, the main cell type affected in Gaucher
disease (Jmoudiak & Futerman, 2005), and to enhance inter-
nalization by mannose receptors on the surfaces of the
macrophages (Brady, 2006), production of Cerezyme
1
involves the sequential deglycosylation of GlcCerase,
using -neuraminidase, -galactosidase and -N-acetyl-
glucosaminidase, to expose terminal mannose residues,
leaving the core glycan, an oligosaccharide which consists of
five sugars, namely two N-acetylglucosamines and three
mannoses.
In our previous structure determinations (Premkumar et al.,
2005; Dvir et al., 2003), Cerezyme
1
was partially deglycosy-
lated prior to crystallization using N-glycosidase F, which
removes carbohydrate chains by cleaving the amide bonds
between Asn residues and N-acetylglucosamine (GlcNAc;
Han & Martinage, 1992), but does not necessarily remove all
carbohydrate chains from native proteins; a similar protocol
was used to obtain another recently reported Cerezyme
1
structure (PDB code 2f61; Liou et al., 2006).
In order to alleviate concerns that partial deglycosylation
might alter its three-dimensional structure, we have now
solved the structure of intact Cerezyme
1
(GCase) obtained
without N-glycosidase F treatment. The crystals display the
same space group as the partially deglycosylated Cerezyme
1
(pDG-GCase); moreover, there are no fundamental differ-
ences between the two structures, although some novel
conformations are observed in the lid region around the active
site (Premkumar et al., 2005). In addition, we demonstrate that
GlcCerase bears a strong structural similarity in its active-site
region to xylanase (Larson et al., 2003), a glycoside hydrolase
that shows the highest sequence similarity to GlcCerase
among structures deposited in the PDB.
2. Experimental procedures
2.1. Crystallization
Cerezyme
1
(GCase; Genzyme Corporation), obtained
from patient leftovers, was dissolved in 100 mMNaCl, 50 mM
2-(4-morpholino)ethanesulfonic acid (MES) pH 5.5 at a
concentration of 1–2 mg ml
1
. The sample was washed with
the same buffer and concentrated to 4–5 mg ml
1
in a
Centricon device using a filter with a cutoff size of 30 kDa.
GCase crystals were obtained by microbatch crystallization
using a Douglas Instruments IMPAX I-5 robot. The crystal-
lization solution contained a 1:1 ratio of the concentrated
enzyme solution and 2 M(NH
4
)
2
SO
4
,0.1MBis-Tris pH 5.5.
Crystallization was performed under oil (D’Arcy et al., 2003)
for 5–14 d at 293 K. Data were collected on beamline ID23eh1
at the ESRF synchrotron facility in Grenoble, France. Crystals
were mounted and flash-cooled at 100 K. X-ray diffraction
images were processed using XDS and XSCALE (Kabsch,
1993). Reflections were converted to a format suitable for
REFMAC5 (Murshudov et al., 1997) using XDSCONV and
processed in CCP4 (Collaborative Computational Project,
Number 4, 1994). Table 1 summarizes data collection and
processing.
2.2. Structure determination and refinement
Initial phases were obtained by molecular replacement
using Phaser (McCoy et al., 2005) software. Molecule Aof
pDG-GCase (PDB code 1ogs; Dvir et al., 2003) was used as a
starting model for molecular replacement. REFMAC5
(Murshudov et al., 1997) was used for refinement and Coot
research papers
Acta Cryst. (2006). D62, 1458–1465 Brumshtein et al. Acid--glucosidase 1459
Figure 1
Reactions catalyzed by GlcCerase and xylanase.
Table 1
Data-collection statistics for GCase.
Values in parentheses are for the highest resolution shell (2.97–2.9 A
˚).
Radiation wavelength (A
˚) 0.9759
Temperature (K) 100
Space group C222
1
Unit-cell parameters (A
˚)a= 108.6, b= 280.8, c= 91.0
Resolution range (A
˚) 40–2.9
No. of observed reflections 228034 (16759)
No. of unique reflections 30978 (2227)
Completeness (%) 98.7 (98.5)
hIi/(I) 13.75 (5.3)
R
mrgd-F
(%) 10.6 (27.0)
electronic reprint
(Emsley & Cowtan, 2004) was used to build the model and fit
it into the electron density. During refinement and model
building, twofold noncrystallographic symmetry was applied,
except for regions where different conformations were clearly
detectable (Kleywegt, 1996). Residues 312–319 (loop 3) in
molecule Bof the GCase model did not fit the electron
density. Consequently, they were omitted from the refinement
and rebuilt into a new conformation.
Electron-density maps revealed sugar residues attached to
some of the Asn residues. Sugars were modelled into electron
density according to the putative oligosaccharide sequences
reported by the Genzyme Corporation for Cerezyme
1
(US
patent Nos. 5236838 and 5549892; Murray, 1987). The con-
formations and angles of the oligosaccharides were confirmed
with PDB-CARE (Lutteke & von der Lieth, 2004) and CARP
(Lutteke et al., 2005). Additional sugars, SO2
4ions and water
molecules were added with improvement of electron-density
maps and phases in subsequent refinement. Refinement
results are summarized in Table 2.
2.3. Comparison of the structures of GCase and xylanase
ABLAST search for sequences in the PDB (Berman et al.,
2000) was performed using the GCase sequence (Horowitz et
al., 1989). The best fit was achieved for xylanase (EC 3.2.1.8;
PDB code 1nof), a member of glycoside hydrolase family 5
(Larson et al., 2003). According to BLASTP (Altschul et al.,
1997), the expectation value is 5 10
3
, with 18% identity and
41% similarity. The crystal structure of xylanase reveals a fold
similar to that of GCase. Therefore, 1nof, which was deter-
mined to 1.42 A
˚resolution (Larson et al., 2003), was used to
cross-validate structural features of GCase (Keen et al., 1996).
2.4. Validation and deposition
The final models and structure factors of GCase were
validated with PROCHECK (Laskowski et al., 1993) and
deposited in the PDB with code 2j25.
2.5. Modelling glycans on pDG-GCase
Glycan modelling was performed on a Silicon Graphics Fuel
workstation using INSIGHTII and DISCOVER software
(Accelrys Inc., San Diego, USA). N-linked glycan structures
were generated using the database of glycosidic linkage
conformations (Wormald et al., 2002) and in vacuo energy
minimization to relieve unfavourable steric interactions. The
Asn–GlcNAc linkage conformations and analysis of the Asn
side-chain conformations were based on the observed range of
crystallographic values (Petrescu et al., 2004). The nomen-
clature for the Asn side-chain torsion angles is
1
=NC
C
—C
,
2
=C
—C
—C
—N
. Figures were produced with
PyMol (http://www.pymol.org).
3. Results
3.1. Crystallization
Previous attempts to obtain Cerezyme
1
crystals without
partial deglycosylation were unsuccessful (Dvir et al., 2003;
Liou et al., 2006; Premkumar et al., 2005). In the current study,
using a much larger screen of crystallization conditions (956
different conditions, excluding optimizations), employing the
microbatch technique under oil (Chayen, 1998) and utilizing a
Douglas Instruments IMPAX I-5 crystallization robot at room
temperature, we were finally able to obtain crystals of Cere-
zyme
1
without prior deglycosylation (GCase). To improve
diffraction, the conditions which produced the best crystals
were chosen for further optimization. GCase crystallizes in the
same space group as previously reported for pDG-GCase, i.e.
C222
1
, and with similar unit-cell parameters.
Analysis of the two structures indicates that all glycosyl-
ation sites are adjacent to empty cavities in the crystal, thus
allowing placement of the sugars in these spaces without
generating steric clashes that would hinder crystallization. The
asymmetric unit of pDG-GCase was shown to contain two
copies of the GlcCerase molecule, molecules Aand B, which,
although very similar, are not completely identical to each
other in conformation (Dvir et al., 2003; Premkumar et al.,
2005). They also differ in the number of sugar molecules for
which electron density can be assigned (Dvir et al., 2003).
Some of the glycoside chains from adjacent asymmetric units
make contacts with each other. These contacts stabilize and
order those oligosaccharide chains in the crystal, thus making
them visible in the electron-density maps.
The structure of GCase was refined to 2.9 A
˚resolution.
Some conformational differences were observed between
molecules Aand Bin GCase and the corresponding molecules
in pDG-GCase (PDB codes 1ogs and 1y7v; Dvir et al., 2003;
Premkumar et al., 2005). The changes seen are mainly in loops
near the active site.
3.2. Glycosylation of GCase
Owing to crystal-packing constraints, the glycans in mole-
cule Bare less mobile than those in molecule A(Fig. 2a). Thus,
in molecule Ba core glycan chain containing five sugar resi-
dues is seen attached to Asn19, three sugars are seen attached
research papers
1460 Brumshtein et al. Acid--glucosidase Acta Cryst. (2006). D62, 1458–1465
Table 2
Refinement statistics for the GCase structure.
Resolution range of refinement (A
˚) 29.6–2.9
R
work
(%) 21.5
R
free
(%) 27.3
Monomers per ASU 2
No. of different NCS groups 2
Average Bfactor (A
˚
2
) 33.6
R.m.s. deviations from ideal values
Bond lengths (A
˚) 0.015
Bond angles () 1.590
Torsion angles () 6.627
Estimated coordinate error
E.s.d.§ from Luzzati plot 0.38
ESU}based on Rfree (A
˚)0.42
ESU based on maximum likelihood (A
˚) 0.31
ESU for Bvalues based on maximum likelihood (A
˚
2
) 17.01
Ramachandran outliers (%) 0.2
R
work
=PjFojjFcj=PjFoj, where F
o
denotes the observed structure-factor
amplitude and F
c
the structure-factor amplitude calculated from the model. R
free
is
for 5% of randomly chosen reflections excluded from the refinement. § E.s.d.,
estimated standard deviation. }ESU, estimated standard uncertainty
electronic reprint
to Asn59, two to Asn146 and none to Asn270 (Table 3). In
molecule Aonly two sugars can be detected on Asn19, one on
Asn146 and none on Asn59 and Asn270. The glycans attached
research papers
Acta Cryst. (2006). D62, 1458–1465 Brumshtein et al. Acid--glucosidase 1461
to residue Asn19-B(i.e. Asn19 of molecule B) and to Asn59-B
of a symmetry-related molecule are in contact with each other
(Fig. 2b), causing the glycan chains to become ordered and, as
a consequence, visible in the electron-density map. The two
sugar moieties on the glycan of Asn146-Bdo not make crystal
contacts, consistent with the low number of sugar residues
visible on Asn146 (Fig. 2c). The fact that we do not detect any
sugars on residue Asn59-Amay be a consequence of the
flexibility of this glycan in the crystal. Residues 59–64 adopt
two different conformations in molecules Aand Bowing to
different crystal contacts.
3.3. Structural characteristics of GCase
Some conformational differences, as distinguished by
differences in and angles (Kleywegt, 1996), were detected
between molecules Aand Bof the asymmetric unit of GCase
(Fig. 3).
3.3.1. Lid at the entrance to the active site. We previously
described two loops in pDG-GCase whose conformations
control access to the active site (Dvir et al., 2003; Premkumar
et al., 2005). In molecule Aof GCase, we now detect confor-
mational differences in an additional loop, loop 3 (residues
312–319; Table 4), that arise in conjunction with changes in the
-sheet structure of residues 341–344, which are located
adjacent to the catalytic residue Glu340 (Fig. 4). The re-
arrangement of loop 3 consists of side-chain flips of both
Trp312 and Trp378 (Fig. 5), which are associated with the
transition of the conformations of residues 312–319 and 341–
344 from a loop and a -strand, respectively, to coils. However,
these differences do not affect the conformation of the cata-
lytic residues, since the distances between the side chains of
Glu235 and Glu340 remain similar (Table 5). Moreover, the
Figure 2
Packing of glycans in the GCase crystal. (a) Glycans in the cavities between the symmetry-related molecules in the crystal. Molecule Bof the asymmetric
unit is shown in blue and the symmetry-related molecules are shown in yellow. The regions labelled Band Care magnified in (b) and (c), respectively. (b)
Magnification of a view of the sugars bound to Asn59 (blue) and to Asn19 on a symmetry-related molecule (yellow). (c) Magnification of a view of the
sugars bound to Asn146 (blue).
Table 3
Number of bound sugar residues observed in the crystal structures of
GlcCerase.
Asn19 Asn59 Asn146 Asn270
2j25-A201 0
2j25-B532 0
1ogs-A100 0
1ogs-B200 0
1y7v-A200 0
1y7v-B200 0
2f61-A100 0
2f61-B200 0
electronic reprint
conformation of His311 (see below) does not change. In
contrast, we detected no new conformations of loops 1–3 in
molecule Brelative to 1ogs and 1y7v, but did detect a novel
combination of previously observed conformations of the
loops.
3.3.2. Active site. Table 5 shows that the distances between
Glu340 and Glu235, the catalytic site residues of GCase, are
5A
˚, which is consistent with a retaining mechanism (Davies
& Henrissat, 1995) of catalytic activity. It is also seen from this
table that the average distances between O
"1
O
"1
and
O
"2
O
"2
are similar in all the crystal structures reported to
date. The differences in distances between individual
carboxylate groups of the catalytic residues of 2j25 and of 1ogs
and 1y7v arise from changes in
2
torsion angles. Since the
acidic pK
a
of the pH-activity profile is around pH 4.7 (Liou et
al., 2006) and the pH values for the crystal structures exam-
ined here lie on either side of this pK
a
value (Dvir et al., 2003;
Premkumar et al., 2005), it is plausible that the observed
differences are related to the protonation states of the active-
site carboxyl groups. Thus, 1ogs and 1y7v, which were both
crystallized at pH 4.6, would be expected to have both
carboxyl groups protonated. This would explain both the
greater inter-residue distances and changes in
2
torsion
angles relative to 2j25. The latter, having been crystallized at
pH 5.5, would be expected to have only one carboxyl group
protonated, which should allow closer approach of the two
carboxylates.
3.3.3. Anion-binding sites. In GCase, there are two clusters
of SO2
4ions, which were previously detected in 1ogs and 1y7v
but not analyzed in detail (Fig. 6a); only one of these clusters
was described in 2f61 (Liou et al., 2006). Each cluster contains
2–3 SO2
4ions. One cluster is located close to residues 12, 44,
45, 353 and 356–358, near the active site (Fig. 6a). The second
is near residues 79, 228, 277 and 306. Since the negative charge
research papers
1462 Brumshtein et al. Acid--glucosidase Acta Cryst. (2006). D62, 1458–1465
Figure 3
A, difference plot between main-chain angles in molecules of the
asymmetric unit of GCase (PDB code 2j25). The differences between
angles are constrained to be between 180and +180.
Table 4
Conformational classification of loops L1, L2 and L3 in the various
GlcCerase crystal structures.
Loop L3, 312–319 L1, 341–347 L2, 393–399
Conformation
Open† 2j25-B2j25-A, 2j25-B2j25-B
1ogs-A, 1ogs-B1ogs-A1ogs-B
1y7v-A, 1y7v-B1y7v-A, 1y7v-B
2f61-A, 2f61-B2f61-A2f61-B
Closed† 2j25-A2j25-A
1ogs-B1ogs-A
1y7v-A, 1y7v-B
2f61-B2f61-A
The assignment of loop conformations as open or closed is based on the criteria
established for 1y7v (Premkumar et al., 2005) and referred to in the discussion.
Figure 4
Conformational classification of loops 1, 2 and 3 (Kleywegt, 1996). The
structures of the loops of 2j25-A(GCase) are in yellow, 2j25-Bin green,
1y7v (pDG-GCase) in magenta, 1ogs-A(pDG-GCase) in orange, 1ogs-B
in red, 2f61-A(pDG-GCase) in black and 2f61-Bin grey. Catalytic
residues are in red.
Table 5
Interatomic distances between side chains of catalytic residues.
Distances are shown in A
˚for the following residues: Glu235–Glu340
(O
"1
O
"1
), Glu235–Glu340 (O
"2
O
"2
) and Glu235–His311 (O
"2
N
1
).
The average distance is between the carboxyl O atoms of the catalytic
glutamates. In the case of 1nof, the distances are measured between residue
Glu165 and residues Glu253 and His230.
O
"1
O
"1
O
"2
O
"2
Average
distance
Closest
approach O
"2
N
1
1ogs-A5.9 4.2 5.05 4.2 2.7
1ogs-B6.2 4.3 5.25 4.3 2.8
1y7v-A6.2 4.0 5.10 4.0 2.8
1y7v-B6.1 3.9 5.00 3.9 2.9
2j25-A4.6 4.2 4.40 4.2 3.2
2j25-B4.9 4.1 4.50 4.1 3.1
2f61-A5.0 5.2 5.10 4.7 3.4
2f61-B6.3 3.9 5.10 3.9 2.6
1nof 5.8† 4.2† 5.00 4.2 2.6
The numbering of O
"1
and O
"2
is arbitrary and the assignment in 1nof was therefore
altered so as to correspond to the numbering in the other structures.
electronic reprint
of SO2
4is similar to that of the negatively charged phospho-
lipids required for optimal GCase activity in vivo (Grace et al.,
1994), we suggest that the SO2
4cluster adjacent to the
active site may be involved in the membrane association of
GlcCerase (Fig. 6b).
3.4. Comparison of the active sites of GCase and xylanase
We compared the structure of GCase with that of xylanase
(PDB code 1nof; Larson et al., 2003), a member of glycoside
hydrolase family 5, which shows the highest sequence identity
to GlcCerase of all structures in the PDB. In order to cross-
validate the positioning and distances of the catalytic residues,
the GCase structure was aligned with that of xylanase (Fig. 7).
Comparison of the active sites was attained by aligning the
catalytic residues: Glu235 and Glu340 in GCase and Glu165
and Glu253 in xylanase. This alignment gave nine identical
and two similar residues within 5 A
˚of the catalytic residues,
which align virtually identically. Only one residue, Tyr313 in
GCase, which corresponds to Tyr232 in xylanase, shows
conformational variability.
Based on the structural and sequence alignment, His311 in
GCase corresponds to His230 in xylanase. The distances
between His311 and Glu235 in GCase and between His230
and Glu165 in xylanase are compatible with hydrogen
bonding. This conserved histidine residue may introduce
additional hydrogen-bond coordination between the catalytic
residues and thus play a role in stabilizing the active sites of
both GCase and xylanase.
4. Discussion
The major conclusion of the current study is that partial
deglycosylation of Cerezyme
1
(pDG-GCase) by N-glycosi-
dase F has no major effect on the conformation of GlcCerase,
thus validating our previous structural analyses of crystals
obtained after N-glycosidase F treatment.
We observe a significant number of Asn-linked glycans in
the GCase structure. The packing of the protein molecules in
the crystal allows the insertion of sugar chains into the free
cavities, which are sufficiently large to accommodate the
glycans without introducing steric clashes with symmetry-
related molecules. We are able to observe two large segments
of two glycan chains, those attached to Asn19-Band Asn59-B,
research papers
Acta Cryst. (2006). D62, 1458–1465 Brumshtein et al. Acid--glucosidase 1463
Figure 6
Putative membrane binding sites in crystal structures of GlcCerase. Catalytic residues are shown in red and loops 1–3 in green. Substrate binding was
modelled according to Dvir et al. (2003) and shown in yellow. (a) Some of the sulfates are shown in space-filling representation in red. (b) Side-on view
[rotated 90around the horizontal axis relative to (a)] of a putative mode of binding of GlcCerase to negatively charged phospholipids.
Figure 5
Conformational changes in Trp312 and Trp378 and movements in the
backbone near the active site. Yellow, 2j25-A; green, 2j25-B; orange, 1ogs-
A; red, 1ogs-B; black, 2f61-A; grey, 2f61-B; magenta, 1y7v; catalytic
residue in red.
electronic reprint
which are in contact with each other and thus mutually
stabilize each other’s conformations. It was previously
reported that the visualization of a sequence of seven sugars of
a glycan chain in Erythrina corallodendron lectin could simi-
larly be ascribed to its immobilization by a symmetry-related
molecule, whereas only chitobioses could be detected on
chains that were not stabilized by such intermolecular inter-
actions (Shaanan et al., 1991).
The observation of sugar residues attached to Asn19 in
pDG-GCase suggested that this residue may be inaccessible to
enzymatic cleavage by N-glycosidase F. This was verified by
molecular modelling, which was performed to determine
whether all four N-linked sites are available for glycosylation
in the monomer and whether the glycans modelled at these
sites block access to the active site. Modelling of the crystal
packing for pDG-GCase showed that except for Asn19, the
putative glycosylation sites could not be occupied with the
observed Asn side-chain conformations. According to the
modelling, a glycan attached to Asn146-Ain pDG-GCase with
the observed side-chain torsion angles (Petrescu et al., 2004)
would clash sterically with a glycan attached to Asn146-B,
indicating that both sites cannot be occupied simultaneously.
Similarly, a core glycan chain modelled on Asn19-Bwould
clash with a modelled glycan on Asn59 of the adjacent B
chain. Thus, the fact that well resolved electron density was
seen for a glycan on Asn19-Bin pDG-GCase implies that
Asn59 must be unoccupied. The presence of glycans on Asn59
and Asn146 in the new GCase structure causes changes in the
1
and
2
torsion angles of the Asn side chains (Petrescu et al.,
2004), thus eliminating steric clashes.
Despite the differences in crystallization conditions and
glycosylation patterns, there are only minor differences
between the pDG-GCase and GCase structures, which have a
similar asymmetric unit and crystal packing, as confirmed by
the low r.m.s. deviations (Table 6). However, several small but
important differences were detected in the conformation of
the lid in molecule Aof GCase, encompassed by loops 1–3
(Premkumar et al., 2005). Importantly, these differences
cannot be ascribed to glycosylation, since molecule Bin the
GCase structure shows conformations of loops 1–3 that are
identical to those seen in pDG-GCase. Molecule Ain GCase
displays some additional changes near the active site relative
to pDG-GCase, in residues Trp312, Trp378 and 341–344. Most
significant are the changes in Trp312 and Trp378, whose side-
chain conformations depend on the conformation of loop 3; as
a result, the conformations of residues 341–344 are also
altered. It should be stressed that even though these loops
control access to the active site, substrate binding must
produce a conformational change in the loops in order to
allow the process to occur without the introduction of steric
clashes, similar to that seen in glycosyl transferases (Qasba et
al., 2005).
Comparison of the catalytic residues Glu235 and Glu340 in
the various GlcCerase structures reveals minor positional
differences, some of which arise from changes in the torsion
angles of Glu235. The side-chain conformation of Glu235
depends on interactions with surrounding residues via
hydrogen bonds. This may explain the differences in the
distances between Glu235 and Glu340 in the various struc-
tures solved to date (Table 5). Moreover, it was recently
implied that the structure of pDG-GCase may not be valid
owing to the loss of activity caused by N-glycosidase F treat-
ment (Liou et al., 2006). That this is not the case is supported
by the similarity in distances between Glu235 and Glu340
(1ogs, 1y7v and 2j25 versus 2f61). Moreover, we have shown
that pDG-GCase is fully catalytically active (Premkumar et al.,
2005). Unfortunately, no data are presented to document the
purported loss of activity produced by N-glycosidase F treat-
ment (Liou et al., 2006). In addition, considerable concern
remains about the lack of correlation between the electron
density and the reported coordinates in some regions of 2f61
(Liou et al., 2006). Irrespective of these concerns, the struc-
tures of 2f61 and of 1ogs, 1y7v and 2j25 are all very similar,
with the minor exceptions noted above with respect to 2j25.
In summary, both the pDG-GCase and GCase structures
reported by us are fully compatible with the catalytic
research papers
1464 Brumshtein et al. Acid--glucosidase Acta Cryst. (2006). D62, 1458–1465
Table 6
Comparison of GlcCerase monomers in various crystal structures.
R.m.s. deviations are shown in A
˚.
1ogs-A1ogs-B1y7v-A1y7v-B2f61-A2f61-B
2j25-A0.35 0.37 0.34 0.34 0.37 0.39
2j25-B0.37 0.34 0.33 0.31 0.39 0.35
2f61-A0.19 0.33 0.30 0.33
2f61-B0.31 0.17 0.30 0.28
1y7v-A0.25 0.26
1y7v-B0.27 0.24
Figure 7
Comparison of the active sites of GCase and xylanase. Residues within a
5A
˚radius of the active site are displayed. Yellow, 2j25-A(GCase); green,
2j25-B; blue, 1nof (xylanase); orange, 1ogs-A(pDG-GCase); red, 1ogs-B;
black, 2f61-A(pDG-GCase); grey, 2f61-B. The first number refers to
residue in GCase and the second number to the corresponding residue in
xylanase.
electronic reprint
mechanism of GlcCerase and provide the molecular tools for
detailed structure–function analysis, which may lead to
improved GlcCerase for use in enzyme-replacement therapy.
This work was supported by the Horowitz Foundation, the
National Gaucher Foundation, the Oxford Glycobiology
Institute, the Yeda fund of the Weizmann Institute, the
Divadol Foundation, the Bruce Rosen Foundation, the
Kalman and Ida Wolens Foundation, the Jean and Jula
Goldwurm Memorial Foundation, the Kimmelman Center for
Biomolecular Structure and Assembly, the Benoziyo Center
for Neuroscience and the Minerva Foundation. We thank the
staff at beamline ID23eh1 at the ESRF synchrotron in
Grenoble for assistance during data collection, Dr Orly Dym
for help in examining the structures and Dr Harry Greenblatt
and Yehudit Hasin for commenting on the manuscript. JLS is
the Morton and Gladys Pickman Professor of Structural
Biology and AHF is the Joseph Meyerhoff Professor of
Biochemistry.
References
Altschul, S. F., Madden, T. L., Schaffer, A. A., Zhang, J., Zhang, Z.,
Miller, W. & Lipman, D. J. (1997). Nucleic Acids Res. 25, 3389–3402.
Berman, H. M., Westbrook, J., Feng, Z., Gilliland, G., Bhat, T. N.,
Weissig, H., Shindyalov, I. N. & Bourne, P. E. (2000). Nucleic Acids
Res. 28, 235–242.
Beutler, E. & Grabowski, G. A. (2001). The Metabolic and Molecular
Bases of Inherited Disease, edited by C. R. Scriver, W. S. Sly, B.
Childs, A. L. Beaudet, D. Valle, K. W. Kinzler & B. Vogelstein, pp.
3635–3668. New York: McGraw–Hill.
Brady, R. O. (2006). Annu. Rev. Med. 57, 283–296.
Bruhn, H. (2005). Biochem. J. 389, 249–257.
Chayen, N. E. (1998). Acta Cryst. D54, 8–15.
Collaborative Computational Project, Number 4 (1994). Acta Cryst.
D50, 760–763.
D’Arcy, A., Mac Sweeney, A., Stihle, M. & Haber, A. (2003). Acta
Cryst. D59, 396–399.
Davies, G. & Henrissat, B. (1995). Structure,3, 853–859.
Dvir, H., Harel, M., McCarthy, A. A., Toker, L., Silman, I., Futerman,
A. H. & Sussman, J. L. (2003). EMBO Rep. 4, 704–709.
Emsley, P. & Cowtan, K. (2004). Acta Cryst. D60, 2126–2132.
Futerman, A. H. & van Meer, G. (2004). Nature Rev. Mol. Cell Biol. 5,
554–565.
Grabowski, G. A., Gatt, S. & Horowitz, M. (1990). Crit. Rev. Biochem.
Mol. Biol. 25, 385–414.
Grace, M. E., Newman, K. M., Scheinker, V., Berg-Fussman, A. &
Grabowski, G. A. (1994). J. Biol. Chem. 269, 2283–2291.
Han, K. K. & Martinage, A. (1992). Int. J. Biochem. 24, 19–28.
Horowitz, M., Wilder, S., Horowitz, Z., Reiner, O., Gelbart, T. &
Beutler, E. (1989). Genomics,4, 87–96.
Jmoudiak, M. & Futerman, A. H. (2005). Br. J. Haematol. 129,
178–188.
Kabsch, W. (1993). J. Appl. Cryst. 26, 795–800.
Keen, N. T., Boyd, C. & Henrissat, B. (1996). Mol. Plant Microbe
Interact. 9, 651–657.
Kleywegt, G. J. (1996). Acta Cryst. D52, 842–857.
Larson, S. B., Day, J., Barba de la Rosa, A. P., Keen, N. T. &
McPherson, A. (2003). Biochemistry,42, 8411–8422.
Laskowski, R. A., Moss, D. S. & Thornton, J. M. (1993). J. Mol. Biol.
231, 1049–1067.
Liou, B., Kazimierczuk, A., Zhang, M., Scott, C. R., Hegde, R. S. &
Grabowski, G. A. (2006). J. Biol. Chem. 281, 4242–4253.
Lutteke, T., Frank, M. & von der Lieth, C. W. (2005). Nucleic Acids
Res. 33, D242–D246.
Lutteke, T. & von der Lieth, C. W. (2004). BMC Bioinformatics,5, 69.
McCoy, A. J., Grosse-Kunstleve, R. W., Storoni, L. C. & Read, R. J.
(2005). Acta Cryst. D61, 458–464.
Murray, G. J. (1987). Methods Enzymol. 149, 25–42.
Murshudov, G. N., Vagin, A. A. & Dodson, E. J. (1997). Acta Cryst.
D53, 240–255.
Petrescu, A. J., Milac, A. L., Petrescu, S. M., Dwek, R. A. & Wormald,
M. R. (2004). Glycobiology,14, 103–114.
Premkumar, L., Sawkar, A. R., Boldin-Adamsky, S., Toker, L.,
Silman, I., Kelly, J. W., Futerman, A. H. & Sussman, J. L. (2005). J.
Biol. Chem. 280, 23815–23819.
Qasba, P. K., Ramakrishnan, B. & Boeggeman, E. (2005). Trends
Biochem. Sci. 30, 53–62.
Sawkar, A. R., Adamski-Werner, S. L., Cheng, W. C., Wong, C. H.,
Beutler, E., Zimmer, K. P. & Kelly, J. W. (2005). Chem. Biol. 12 ,
1235–1244.
Shaanan, B., Lis, H. & Sharon, N. (1991). Science,254, 862–866.
Vaccaro, A. M., Tatti, M., Ciaffoni, F., Salvioli, R., Barca, A. & Scerch,
C. (1997). J. Biol. Chem. 272, 16862–16867.
Wormald, M. R., Petrescu, A. J., Pao, Y. L., Glithero, A., Elliott, T. &
Dwek, R. A. (2002). Chem. Rev. 102, 371–386.
research papers
Acta Cryst. (2006). D62, 1458–1465 Brumshtein et al. Acid--glucosidase 1465
electronic reprint
... The refined structure shows one molecule of GCase bound to one molecule of Nb1 in the asymmetric unit ( Figure 3A). As previously described, the structure of GCase displays a globular fold formed by 3 noncontiguous domains: domain I (residues 1-29 and 383-414) is a small three-stranded antiparallel β-sheet, domain II (residues 30-77 and 431-497) forms an eight-stranded β-barrel, and domain III (residues 78-382 and 415-430) adopts a (β/α)8 triose-phosphate isomerase (TIM) barrel, which contains the active site and the two catalytic residues E235 and E340 39,[49][50][51] . During refinement, sugar moieties were added on residues N19 and N270, which are well-known and characterized glycosylation sites 39 . ...
Preprint
Full-text available
The enzyme glucocerebrosidase (GCase) catalyses the hydrolysis of glucosylceramide to glucose and ceramide within lysosomes. Homozygous or compound heterozygous mutations in the GCase-encoding GBA1 gene cause the lysosomal storage disorder Gaucher disease, while heterozygous mutations are the most frequent genetic risk factor for Parkinson's disease. These mutations commonly affect GCase stability, trafficking or activity. Here, we report the development and characterization of nanobodies (Nbs) targeting and acting as chaperones for GCase. We identified several Nb families that bind with nanomolar affinity to GCase. Based on biochemical characterization, we grouped the Nbs in two classes: Nbs that improve the activity of the enzyme and Nbs that increase GCase stability in vitro. A selection of the most promising Nbs was shown to improve GCase function in cell models and positively impact the activity of the N370S mutant GCase. These results lay the foundation for the development of new therapeutic routes.
... Recently, it has been found that a transglucosylation reaction occurs for the lysosomal acid GCase to synthesize Glc-cholesterol and Gal-cholesterol [91]. The five N-glycans of GCase control its structure [92]. Glucocerebrosidase II (GCase II, GBA2) is a non-lysosomal enzyme found in the ER, Golgi and plasma membrane. ...
Article
Full-text available
Glycosphingolipids (GSLs) are a specialized class of membrane lipids composed of a ceramide backbone and a carbohydrate-rich head group. GSLs populate lipid rafts of the cell membrane of eukaryotic cells, and serve important cellular functions including control of cell-cell signaling, signal transduction and cell recognition. Of the hundreds of unique GSL structures, anionic gangliosides are the most heavily implicated in the pathogenesis of lysosomal storage diseases (LSDs) such as Tay-Sachs and Sandhoff disease. Each LSD is characterized by the accumulation of GSLs in the lysosomes of neurons, which negatively interact with other intracellular molecules to culminate in cell death. In this review, we summarize the biosynthesis and degradation pathways of GSLs, discuss how aberrant GSL metabolism contributes to key features of LSD pathophysiology, draw parallels between LSDs and neurodegenerative proteinopathies such as Alzheimer's and Parkinson's disease and lastly, discuss possible therapies for patients.
... DNJ derivatives 10 and 12a-d were examined for the correction capacity of the CFTR function in CF-KM4 cells [93] using iodide efflux experiments ( Figure 5) [94]. Iminosugars 13-15 bearing saturated alkyl chains were also considered [8,[95][96][97]. No significant rescue capacity of F508del-CFTR activity was observed for most iminosugars, including those bearing longer saturated alkyl chains (13)(14)(15) than NBDNJ, as well as those with alkyl chains having fluorine atoms (12a-d). ...
Article
Full-text available
Iminosugars are sugar analogues endowed with a high pharmacological potential. The wide range of biological activities exhibited by these glycomimetics associated with their excellent drug profile make them attractive therapeutic candidates for several medical interventions. The ability of iminosugars to act as inhibitors or enhancers of carbohydrate-processing enzymes suggests their potential use as therapeutics for the treatment of cystic fibrosis (CF). Herein we review the most relevant advances in the field, paying attention to both the chemical synthesis of the iminosugars and their biological evaluations, resulting from in vitro and in vivo assays. Starting from the example of the marketed drug NBDNJ (N-butyl deoxynojirimycin), a variety of iminosugars have exhibited the capacity to rescue the trafficking of F508del-CFTR (deletion of F508 residue in the CF transmembrane conductance regulator), either alone or in combination with other correctors. Interesting results have also been obtained when iminosugars were considered as anti-inflammatory agents in CF lung disease. The data herein reported demonstrate that iminosugars hold considerable potential to be applied for both therapeutic purposes.
... GBA binding and subsequent internalization into target cell macrophages is mediated by interactions between terminal mannose residues on GBA and mannose receptors on the cell surface [12,13]. GBA is known to contain five potential N-glycosylation sites [14,15]. Exposure of terminal mannose residues by a sequential in vitro deglycosylation strategy in Cerezyme ® production dramatically improved its uptake into target cells [13,[16][17][18]. ...
Article
Gaucher disease is an inherited metabolic disease caused by genetic acid β -glucosidase (GBA) deficiency and is currently treated by enzyme replacement therapy. For uptake into macrophages, GBA needs to carry terminal mannose residues on their N-glycans. Knockout mutant rice of N-acetylglucosaminyltransferase-I (gnt1) have a disrupted N-glycan processing pathway and produce only glycoproteins with high mannose residues. In this study, we introduced a gene encoding recombinant human GBA into both wild-type rice (WT) and rice gnt1 calli. Target gene integration and mRNA expression were confirmed by genomic DNA PCR and Northern blotting, respectively. Secreted rhGBAs in culture media from cell lines originating from both WT (WT-GBA) and rice gnt1 (gnt1-GBA) were detected by Western blotting. Each rhGBA was purified by affinity and ion exchange chromatography. In vitro catalytic activity of purified rhGBA was comparable to commercial Chinese hamster ovary cell-derived rhGBA. N-glycans were isolated from WT-GBA and gnt1-GBA and analyzed by using matrix-assisted laser desorption/ionization time-of-flight mass spectrometry. The amounts of high mannose-type N-glycans were highly elevated in gnt1-GBA (100%) compared to WT-GBA (1%).
... The orientation of both proteins anchored to the membrane was consistent with the experiments. 29,73,117 GCase was oriented with the loops at the entrance of the binding site facing the phospholipid membrane. As GluCer anchored in the membrane, it was anticipated that the active site will be facing the membrane. ...
Conference Paper
Gaucher’s Disease (GD) is a rare recessive disorder produced by the dysfunction of the lysosomal enzyme Glucocerebrosidase (GCase). GCase catalyses the cleavage of the glycolipid Glucosylceramide. The lack of functional GCase leads to the accumulation of its lipid substrate in lysosomes causing GD. GD presents a great phenotypic variation, symptoms ranging from asymptomatic adults to early childhood death due to neurological damage. More than 250 mutations in the protein GCase have been discovered that result in GD. Being able to link structural modifications of each mutation to the phenotypic variation of GD would enhance the understanding of the disease. The aim of this work is to understand the structural dynamics of wild type and mutant GCase. A model of the complex of the enzyme GCase with its facilitator protein, Saposin-C (Sap-C) was generated using Protein-Protein docking (PPD). In this work, a knowledge-based docking protocol that considers experimental data of protein- protein binding has been carried out. Here, a reliable model of the enzyme GCase with its facilitator protein is presented and is consistent with the experimental data. To understand the structural mechanism of function of the enzyme GCase, it was imperative to study its structural dynamics and conformational changes influenced by its interaction with other components including lipid bilayer, facilitator protein or substrate. Coarse-Grained MD (CG-MD) was employed to study lipid self-assembly and membrane insertion of the complex. Classical Atomistic MD (AT-MD) was used to study the dynamics of the interactions between different components of the simulation. Furthermore, the results of ten different AT-MD simulations sampling 9 s have been analysed. An activation method of GCase by Sap-C has been proposed, the change in conformation of GCase when its facilitator protein is present has been highlighted, through the stabilization of the loops at the entrance of the binding site. The differences in protein-protein binding when GCase is mutated have also been emphasised. Finally, Anharmonic Conformational Analysis and Markov State Models have been used to build a kinetic model of the system. This model supports our activation mechanism hyphothesis.
... Post-translational modifications are vital for active and stable GCase (30,63,64). While protein with these modifications can be prepared from human cell lines, introducing the necessary isotopes for NMR analysis during expression is prohibitively expensive. ...
Preprint
Full-text available
Aggregation of the protein alpha-synuclein (αSyn) into amyloid fibrils is associated with Parkinson's disease (PD), a process accelerated by lipids. Recently, the lysosomal protein glucocerebrosidase (GCase) has been identified as a major risk factor in both genetic and sporadic PD. Here, we use solution state NMR to reveal that GCase directly inhibits lipid induced αSyn amyloidogenesis. Structurally, we show that the mechanism for this requires competition between lipids and GCase for αSyn, binding the N and C termini respectively. The affinity of GCase for the C-terminus of αSyn is such that not only does it inhibit lipid induced amyloid formation, but also it destabilizes mature αSyn amyloid fibrils. These results reveal a competitive molecular "tug-of-war" for αSyn termini by GCase and lipid, providing a mechanistic link between the clinically observed links between changes in GCase abundance and Parkinsons disease.
Chapter
Enzymes are key biological macromolecules that support life by accelerating the conversion of target molecules to desired products in many biochemical reactions. Enzymes are characterized by high affinity, specificity and great catalytic efficiency. Owing to their unique characteristics, enzymes have attracted significant attention for use in therapeutic settings as a distinct class of drugs different from other types of medicines. Enzyme-based therapies are currently in use for the treatment of a wide range of diseases, including leukemia, metabolic disorders, inflammation and cardiovascular disease. However, several challenges, such as immunogenicity and stability, remain. X-ray crystallography has provided key structural insights into the understanding of the molecular basis of diseases and development of enzyme-based therapies. Here, the role of X-ray crystallography in the development of therapeutic enzymes is examined and several examples are provided.
Article
Small-molecule- enzyme enhancement therapeutics (EETs) have emerged as attractive agents for the treatment of lysosomal storage diseases (LSDs), a broad group of genetic diseases caused by mutations in genes encoding lysosomal enzymes, or proteins required for lysosomal function. The underlying enzyme deficiencies characterizing LSDs cause a block in the stepwise degradation of complex macromolecules (e.g. glycosaminoglycans, glycolipids and others), such that undegraded or partially degraded substrates progressively accumulate in lysosomal and non-lysosomal compartments, a process leading to multisystem pathology via primary and secondary mechanisms. Missense mutations underlie many of the LSDs; the resultant mutant variant enzyme hydrolase is often impaired in its folding and maturation making it subject to rapid disposal by endoplasmic reticulum (ER)-associated degradation (ERAD). Enzyme deficiency in the lysosome is the result, even though the mutant enzyme may retain significant catalytic functioning. Small molecule modulators - pharmacological chaperones (PCs), or proteostasis regulators (PRs) are being identified through library screens and computational tools, as they may offer a less costly approach than enzyme replacement therapy (ERT) for LSDs, and potentially treat neuronal forms of the diseases. PCs, capable of directly stabilizing the mutant protein, and PRs, which act on other cellular elements to enhance protein maturation, both allow a proportion of the synthesized variant protein to reach the lysosome and function. Proof-of-principle for PCs and PRs as therapeutic agents has been demonstrated for several LSDs, yet definitive data of their efficacy in disease models and/or in downstream clinical studies in many cases has yet to be achieved. Basic research to understand the cellular consequences of protein misfolding such as perturbed organellar crosstalk, redox status, and calcium balance is needed. Likewise, an elucidation of the early in cellulo pathogenic events underlying LSDs is vital and may lead to the discovery of new small molecule modulators and/or to other therapeutic approaches for driving proteostasis toward protein rescue.
Article
Full-text available
Structure/function relationships of acid beta-glucosidase, the enzyme deficient in Gaucher disease, were evaluated by characterizing the proteins expressed from cDNAs encoding normal and mutant enzymes. Twenty-two Gaucher disease mutations or created mutations were expressed in Spodoptera frugiperda (Sf9) cells and analyzed for catalytic properties, stability, inhibitor binding, and modifier interactions. Many Gaucher disease mutations encoded highly disruptive amino acid substitutions (e.g. P289L and D409V) and produced severely compromised proteins with very reduced activity (kcat < 1% of normal) and/or stability. Six mutant enzymes had sufficient catalytic activity (kcat approximately 5-30% of normal) for extensive studies. The highly conservative substitutions, i.e. F216Y or S364T and V394L, led to severe, but selective, abnormalities of enzyme stability or large decreases in catalytic activity, respectively. The T323I, N370S, and V394L enzymes interacted abnormally with active site-directed inhibitors and localized these residues to the glycon binding region. Selected mutant enzymes were poorly activated by phosphatidylserine (V394L, L444P, and R463C) or by saposin C (L444P and T323I), indicating that the enzyme sites for interaction with these activators were within the carboxyl one-third of the enzyme. Substitutions of Ser, Glu, and/or Gly at residues Asp-443 and/or Asp-445 demonstrated important steric roles for these residues in the active site, but neither is the catalytic nucleophile. Together with previous studies, the present analyses provide an insight into the pathogenesis of Gaucher disease and the functional organization of acid beta-glucosidase.
Article
The Protein Data Bank (PDB; http://www.rcsb.org/pdb/ ) is the single worldwide archive of structural data of biological macromolecules. This paper describes the goals of the PDB, the systems in place for data deposition and access, how to obtain further information, and near-term plans for the future development of the resource.
Article
1. The role played by the modification of protein in determining its fate is reported by us. 2. Post-translational modifications such as acetylation, phosphorylation, sulfation, methylation, hydroxylation, ADP-ribosylation, maturation, amidation, carboxylation, adenylylation, glycosylation, ubiquitination, and prenylation are extensively reviewed. 3. Each post-translational modification's significance and its role played in biological function(s) is summarized in the general discussion and the conclusion's remark is directed at the problems left to solve (e.g. post-translational modification reactions in recombinant protein in modern genetic engineering).
Article
The three-dimensional structure of the lactose complex of the Erythrina corallodendron lectin (EcorL), a dimer of N-glycosylated subunits, was determined crystallographically and refined at 2.0 angstrom resolution to an R value of 0.19. The tertiary structure of the subunit is similar to that of other legume lectins, but interference by the bulky N-linked heptasaccharide, which is exceptionally well ordered in the crystal, forces the EcorL dimer into a drastically different quaternary structure. Only the galactose moiety of the lactose ligand resides within the combining site. The galactose moiety is oriented differently from ligands in the mannose-glucose specific legume lectins and is held by hydrophobic interactions with Ala88, Tyr106, Phe131, and Ala218 and by seven hydrogen bonds, four of which are to the conserved Asp89, Asn133, and NH of Gly107. The specificity of legume lectins toward the different C-4 epimers appears to be associated with extensive variations in the outline of the variable parts of the binding sites.
Article
Human lysosomal beta-glucosidase (D-glucosyl-acylsphingosine glucohydrolase, EC 3.2.1.45) is a membrane-associated enzyme that cleaves the beta-glucosidic linkage of glucosylceramide (glucocerebroside), its natural substrate, as well as synthetic beta-glucosides. Experiments with cultured cells suggest that in vivo this glycoprotein requires interaction with negatively charged lipids and a small acidic protein, SAP-2, for optimal glucosylceramide hydrolytic rates. In vitro, detergents (Triton X-100 or bile acids) or negatively charged ganglioside or phospholipids and one of several "activator proteins" increase hydrolytic rate of lipid and water-soluble substrates. Using such in vitro assay systems and active site-directed covalent inhibitors, kinetic and structural properties of the active site have been elucidated. The defective activity of this enzyme leads to the variants of Gaucher disease, the most prevalent lysosomal storage disease. The nonneuronopathic (type 1) and neuronopathic (types 2 and 3) variants of this inherited (autosomal recessive) disease but panethnic, but type 1 is most prevalent in the Ashkenazi Jewish population. Several missense mutations, identified in the structural gene for lysosomal beta-glucosidase from Gaucher disease patients, are presumably casual to the specifically altered posttranslational oligosaccharide processing or stability of the enzyme as well as the altered in vitro kinetic properties of the residual enzyme from patient tissues.
Article
We report the sequence of the entire human gene encoding beta-glucocerebrosidase and that of the associated pseudogene. The gene contains 11 exons extending from base pair 355 to base pair 7232 in the overall sequence. The gene promoter contains TATA- and CAT-like boxes upstream of the major 5' end of the glucocerebrosidase RNA. The two TATA boxes lie between nucleotides (-23)-(-27) and (-33)-(-39) and the two possible CAT boxes reside between nucleotides (-90)-(-94) and (-96)-(-99) in relation to the major 5' end of the mRNA. The functionality of the promoter region was monitored by coupling it to the bacterial gene coding for chloramphenicol acetyltransferase (CAT) and assaying the expression of the enzyme in cells transfected with this vector. The glucocerebrosidase promoter not only directs synthesis of the bacterial enzyme but also exhibits the same pattern of tissue-specific expression as that of the endogenous gene. An apparently tightly linked pseudogene is approximately 96% homologous to the functional gene. However, introns 2, 4, 6, and 7 have large "deletions" consisting of Alu sequences 313, 626, 320, and 277 bp in length, respectively. It is entirely possible that the ancestral gene lacks these sequences and that they have been inserted into the introns of the functioning gene. There is also a 55-bp deletion from a part of exon 9 flanked by a short inverted repeat. The sequence data should facilitate development of methods for diagnosis of Gaucher disease at the molecular level.
Article
The principles and methods used for enzymatic modification of the carbohydrate portion of glucocerebrosidase are similar to those performed by Ashwell and Morell, Stahl, and others. It is difficult to explain the lack of uptake of native enzyme through binding of the high-mannose type glycopeptide to Man/GlcNAc receptors since approximately 20% of the total oligosaccharides on the native enzyme are high mannose type. Possibly a requirement for multiple sites of attachment to the receptor is not met by a single high-mannose type oligosaccharide per molecule. Alternatively, the presence of complex type oligosaccharides on this enzyme, demonstrated by structural studies, may mask the mannose site and thus account for the poor uptake of native enzyme.
Article
The main-chain bond lengths and bond angles of protein structures are analysed as a function of resolution. Neither the means nor standard deviations of these parameters show any correlation with resolution over the resolution range investigated. This is as might be expected as bond lengths and bond angles are likely to be heavily influenced by the geometrical restraints applied during structure refinement. The size of this influence is then investigated by performing an analysis of variance on the mean values across the five most commonly used refinement methods. The differences in means are found to be highly statistically significant, suggesting that the different target values used by the different methods leave their imprint on the structures they refine. This has implications concerning the actual target values used during refinement and stresses the importance of the values being not only accurate but also consistent from one refinement method to another.